• 検索結果がありません。

Wong’s Equations and Charged Relativistic Particles in Non-Commutative Space

N/A
N/A
Protected

Academic year: 2022

シェア "Wong’s Equations and Charged Relativistic Particles in Non-Commutative Space"

Copied!
21
0
0

読み込み中.... (全文を見る)

全文

(1)

Wong’s Equations and Charged Relativistic Particles in Non-Commutative Space

?

Herbert BALASIN , Daniel N. BLASCHKE, Fran¸cois GIERES § and Manfred SCHWEDA

Institute for Theoretical Physics, Vienna University of Technology, Wiedner Hauptstraße 8-10, A-1040 Vienna, Austria

E-mail: hbalasin@tph.tuwien.ac.at, mschweda@tph.tuwien.ac.at

Los Alamos National Laboratory, Theory Division, Los Alamos, NM, 87545, USA E-mail: dblaschke@lanl.gov

§ Universit´e de Lyon, Universit´e Claude Bernard Lyon 1 and CNRS/IN2P3, Institut de Physique Nucl´eaire, Bat. P. Dirac,

4 rue Enrico Fermi, F-69622-Villeurbanne, France E-mail: gieres@ipnl.in2p3.fr

Received March 02, 2014, in final form October 17, 2014; Published online October 24, 2014 http://dx.doi.org/10.3842/SIGMA.2014.099

Abstract. In analogy to Wong’s equations describing the motion of a charged relativistic point particle in the presence of an external Yang–Mills field, we discuss the motion of such a particle in non-commutative space subject to an externalU?(1) gauge field. We conclude that the latter equations are only consistent in the case of a constant field strength. This formulation, which is based on an action written in Moyal space, provides a coarser level of description than full QED on non-commutative space. The results are compared with those obtained from the different Hamiltonian approaches. Furthermore, a continuum version for Wong’s equations and for the motion of a particle in non-commutative space is derived.

Key words: non-commutative geometry; gauge field theories; Lagrangian and Hamiltonian formalism; symmetries and conservation laws

2010 Mathematics Subject Classification: 81T13; 81T75; 70S05

1 Introduction

Over the last twenty years, non-commuting spatial coordinates have appeared in various contexts in the framework of quantum gravity and superstring theories. This fact contributed to the motivation for studying classical and quantum theories with a finite or infinite number of degrees of freedom on non-commutative spaces. Different mathematical approaches have been pursued and various physical applications have been explored, e.g. see references [8,11,24,35,39] for some partial reviews. Beyond the applications, classical and quantum mechanics on non-commutative space are of interest as toy models for field theories which are more difficult to handle, in particular in the case of interactions with gauge fields. In this respect, we recall a similar situation concerning the coupling of matter to Yang–Mills fields on ordinary space: A coarser level of description for the latter theories has been proposed by Wong [45] who considered the motion of charged point particles in an external gauge field [6, 7, 9, 15, 16, 26, 27, 44].

The latter equations allow for various physical applications, e.g. to the dynamics of quarks and their interaction with gluons [26,27]. Somewhat similar equations, known as Mathisson–

Papapetrou–Dixon equations [13, 30,33] appear in general relativity for a spinning particle in

?This paper is a contribution to the Special Issue on Deformations of Space-Time and its Symmetries. The full collection is available athttp://www.emis.de/journals/SIGMA/space-time.html

(2)

curved space. In this spirit, we will consider in the present work the dynamics of a relativistic

“point” particle in non-commutative space subject to an external U?(1) gauge field (thereby implementing a suggestion made in an earlier work [38], see also [20] for some related studies based on a first order expansion in the non-commutativity parameter).

More precisely, we are motivated by the motion in Moyal space, the latter space having been widely discussed as the arena for field theory in non-commutative spaces. By proceeding along the lines of Wong’s equations (which are discussed in Section 2), we derive a set of equations for the dynamics of the particle in Moyal space in Section3. For the description of the coupling of a particle to a gauge field, the relativistic setting is the most natural one, but our discus- sions (of Wong’s equations in commutative space or of a particle in non-commutative space) could equally well be done within the non-relativistic setting. The particular case of a con- stant field strength is the most tractable one (and the only consistent one for Wong’s equations in non-commutative space) and will be considered in some detail in Section 4. Subsequently in Section5, we briefly recall the different Hamiltonian approaches which have previously been pursued for the formulation of classical mechanics in non-commutative space and we compare the resulting equations governing the dynamics of particles coupled to an electromagnetic field.

In the appendix we present a continuum formulation of Wong’s equations on a generic space-time manifold, a formulation which readily generalizes to Moyal space.

For the motion of a relativistic point particle in four-dimensional (commutative or non- commutative) space-time, the following notations will be used. The metric tensor is given by

µν)µ,ν∈{0,1,2,3} = diag(1,−1,−1,−1)

and we choose the natural system of units (c ≡ 1 ≡~). The proper time τ for the particle is defined (up to an additive constant) by

2=ds2, with ds2≡dxµdxµ= (dt)2−(d~x)2, (1.1) and for the massive particle we haveds2 >0. From (1.1) it follows that ˙x2= 1 where ˙x2 ≡x˙µµ

and ˙xµ≡dxµ/dτ.

2 Reminder on particles in commutative space

Abelian gauge f ield in f lat space. We consider the interaction of a charged massive rela- tivistic particle with an external electromagnetic field given by the U(1) gauge potential (Aµ).

The motion of this particle along its space-time trajectoryτ 7→xµ(τ) is described by the action1 S[x] =−m

Z

ds−q Z

dxµAµ(x(τ)) =−m Z

˙ x2−q

Z

dτx˙µAµ(x(τ)). (2.1) Here,qdenotes the conserved electric charge of the particle associated with the conserved current density (jµ):

jµ(y) =q Z

dτx˙µ(τ)δ4(y−x(τ)), ∂µjµ= 0, Z

d3yj0(y) =q. (2.2) We note that the interaction term in the functional (2.1) may be rewritten in terms of the above current according to

q Z

dτx˙µAµ(x(τ)) = Z

d4yjµ(y)Aµ(y). (2.3)

1To be more precise, in the integral (2.1) the variableτ is viewed as a purely mathematical parameter which is only identified with proper time after deriving the equations of motion from the action. Thus, the relation ˙x2= 1 is only to be used at the latter stage.

(3)

This expression is invariant under infinitesimal gauge transformations, i.e. δλAµ=∂µλ, thanks to the conservation of the current:

δλ Z

d4yjµAµ= Z

d4yjµµλ=− Z

d4y(∂µjµ)λ= 0.

We note that the parameter q which describes the coupling of the particle to the gauge field might in principle depend on the world line parameterτ: if this assumption is made,q(τ) appears under the τ-integral in (2.2), the current jµ is no longer conserved and the coupling (2.3) is no longer gauge invariant. Thus, q is necessarily constant along the path.

Variation of the action (2.1) with respect to xµ and substitution of the relation ˙x2= 1 leads to the familiar equation of motion

mx¨µ=qFµνν, with Fµν ≡∂µAν−∂νAµ. As is well known, the free particle LagrangianLfree=−m√

˙

x2 which represents the first term of the action (2.1) and which is non-linear in ˙x2can be replaced by ˜Lfree = m22 which is linear in ˙x2 since both Lagrangians yield the same equation of motion. Indeed, we will consider the latter Lagrangian in equation (2.12) and in Section5where we go over to the Hamiltonian formulation.

The aim of this work is to generalize this setting to a non-commutative space-time. Since a gauge field on non-commutative space entails a non-Abelian structure for the field strength tensorFµν due to the star product, it is worthwhile to understand first the coupling of a spinless particle to a non-Abelian gauge field on commutative space-time.

Non-Abelian gauge f ield in f lat space. We consider a compact Lie groupG (e.g. G= SU(N) for concreteness) with generators Ta satisfying

Ta, Tb

= ifabcTc and Tr TaTb

ab.

Just as for the Abelian gauge field, the source jµa(x) of the non-Abelian gauge field Aµ(x) ≡ Aaµ(x)Ta (e.g. a field theoretic expression like ¯ψγµaψ involving a multipletψ of spinor fields) is considered to be given by the current density jµa(x) of a relativistic point particle [6,7,15,16, 26,27,44,45]. Instead of an electric charge, the particle moving in an external Yang–Mills field is thus assumed to carry a color-charge or isotopic spin ~q ≡ (qa)a=1,...,dimG which transforms under the adjoint representation of the structure group G. Henceforth, one considers the Lie algebra-valued variableq(τ)≡qa(τ)Tawhich is assumed to beτ-dependent. The particle is then described in terms of its space-time coordinatesxµ(τ) and its isotopic spinq(τ), i.e. it is referred to with respect to geometric space and to internal space. For the moment, we assume q(τ) to represent a given non-dynamical (auxiliary) variable and we will comment on a different point of view below. Its coupling to an external non-Abelian gauge field (Aµ) is now described by the action

S[x] =−m Z

ds− Z

dτx˙µTr{q(τ)Aµ(x(τ))}

=−m Z

˙ x2

Z

dτx˙µqa(τ)Aaµ(x(τ)). (2.4)

The current density (2.2) presently generalizes to a Lie algebra-valued expression jµ ≡ jµaTa given by

jµ(y) = Z

dτx˙µ(τ)q(τ)δ4(y−x(τ)), (2.5)

and thereby the interaction term in the functional (2.4) can be rewritten as

− Z

d4yTr(jµAµ). (2.6)

(4)

For an infinitesimal gauge transformation with Lie algebra-valued parameter λ, i.e. δλAµ = Dµλ≡∂µλ−ig[Aµ, λ], we have

δλ Z

d4yTr{jµAµ}= Z

d4yTr{jµDµλ}=− Z

d4yTr{(Dµjµ)λ}. (2.7) Thus, gauge invariance of the action (2.4) requires the current to be covariantly conserved, i.e. Dµjµ= 0. From (2.5) we can deduce by a short calculation that

(Dµjµ)a(y) = Z

dτDqa

dτ δ4(y−x(τ)), with Dqa dτ ≡ dqa

dτ −igx˙µ[Aµ(x(τ)), q]a, (2.8) hence jµ is covariantly conserved if the charge q is covariantly constant along the world line:

Dqa/dτ = 0 (subsidiary condition).

Variation of the action with respect toxµ(and use of ˙x2 = 1) yields the equations of motion mx¨µ= Tr(qFµν) ˙xν, where Dqa

dτ = 0, (2.9)

and Fµν ≡ ∂µAν −∂νAµ−ig[Aµ, Aν]. By construction, these equations are invariant under infinitesimal gauge transformations parametrized by λ(x) since the chargeqis a non-dynamical (auxiliary) variable transforming as δλq(τ)≡ −ig[q(τ), λ(x(τ))] = 0 for all τ, hence

δλ Tr(qFµν) ˙xν

=−igTr(q[Fµν, λ]) ˙xν = igTr([q, λ]Fµν) ˙xν = 0, δλ

Dq dτ

=−igx˙µλAµ, q] =−ig Dλ

dτ , q

=−igD

dτ[λ, q] = 0.

Equations (2.9), which represent the Lorentz–Yang–Mills force equation for a relativistic particle in an external Yang–Mills field, are known asWong’s equations [45]. More specifically, the rela- tionDqa/dτ = 0 may be viewed as charge transport equation and it is the geometrically natural generalization (to the charge vector (qa)) of the constancy of the chargeq in electrodynamics.

A few remarks are in order here. We should point out that the equation of motion forxµhad already been obtained earlier in curved space by Kerner [25], but the charge transport equation was not established in that setting. We refer to the works [6, 7, 26, 27] for a treatment of dynamics involving the Lagrangian−14R

d3xTr(FµνFµν) of the gauge field: the latter approach yields the covariant conservation law Dµjµ = 0 as a consequence of the equation of motion DνFνµ =jµ and the relation [Dµ, Dν]Fνµ = 0. For a general discussion of the issue of gauge invariance for the coupling of gauge fields to non-dynamical external sources, we refer to [34].

Equations (2.9) and their classical solutions have been investigated in the literature and applied for instance to the study of the quark gluon plasma [26]. The particular case of a constant field strength Fµν (“uniform fields”) exhibits interesting features [9] which will be commented upon in Section 4.

If one regards (qa) as a dynamical variable which satisfies the equation of motionDqa/dτ = 0 and which transforms under gauge variations with the adjoint representation, i.e.

δλq(τ) =−ig[q(τ), λ(x(τ))], (2.10) then equations (2.9) are obviously gauge invariant since Fµν also transforms with the adjoint representation2. However, we emphasize that we started from the action (2.4) to obtain the equation of motion ofxµ, the one ofq following from the requirement of gauge invariance of the

2Ifjµis assumed to transform covariantly, then gauge invariance of the action (2.6) impliesµjµ= 0 as has already been noticed in reference [36].

(5)

initial action3. An action which yields the equations (2.9) as equations of motion of both xµ andq has been constructed in the non-relativistic setting in reference [28]. The relativistic gene- ralization of this approach proceeds as follows. (For simplicity, we put the coupling constant g equal to one.) One introduces a Lie algebra-valued variable Λ(τ) ≡ Λa(τ)Ta where the func- tions Λa(τ) are Grassmann odd, the charge q being defined as an expression which is bilinear in Λ (i.e. a description which is familiar for the spin):

q ≡ −1

2[Λ,Λ]+, i.e. qa=−i

2fabcΛbΛc. (2.11)

The Lagrangian

L(x,x,˙ Λ,Λ)˙ ≡ m 2x˙2+ i

2Tr

ΛDΛ dτ

= m 2x˙2+ i

2Tr(Λ ˙Λ) + ˙xµTr(qAµ), (2.12) is invariant under gauge transformations for which δλAµ=Dµλand

δλΛ(τ) =−i[Λ(τ), λ(x(τ))],

which implies the transformation law (2.10). Moreover, the Lagrangian leads to the following equations of motion for xµ and Λa:

mx¨µ= ˙xνTr{q(∂µAν−∂νAµ)}+ Tr( ˙qAµ), 0 = DΛa

dτ . (2.13)

From (2.11) it follows that Dq =− Λ,

+, hence (2.13) implies that Dqa = 0, i.e. the chargeq is covariantly conserved. Substitution of the latter result into the first of equations (2.13) yields the equation of motion m¨xµ = Tr(qFµν) ˙xν. We note that the Hamiltonian associated to the Lagrangian (2.12) reads

H = 1

2m[pµ−Tr(qAµ)][pµ−Tr(qAµ)],

orH = m22 if expressed in terms of the velocity. The Poisson brackets {xµ, pν}=δµν, {Λab}=−iδab,

which imply the non-Abelian algebra of charges {qa, qb}=fabcqc,

again allow us to recover all previous equations of motion from the evolution equation of dynami- cal variablesF, i.e. from ˙F ={F, H}. In terms of the kinematical momentum Πµ≡pµ−Tr(qAµ), the Hamiltonian readsH = 2m1 Π2 and the Poisson brackets take the form

{xµν}=δµν, {Πµν}= Tr(qFµν), {qa, qb}=fabcqc.

The dynamical variable Λ which allowed for the Lagrangian formulation is well hidden in the latter equations.

Continuum formulation of the dynamics. The field strengthFµν manifests itself physi- cally by the force field, i.e. by an exchange of energy and momentum between the charge carrier

3In fact, if the chargeq is treated as a dynamical variable in the action (2.4), then it amounts to a Lagrange multiplier leading to the equation of motion ˙xµAaµ(x(τ)) = 0 which is not gauge invariant. We thank the anonymous referee for drawing our attention to this point.

(6)

and the field [40]. In the framework of field theory, the physical entities are described by lo- cal fields, i.e. one has a continuum formulation. In order to obtain such a formulation for the particle’s equations of motion (2.9), we have to integrate these relations over the variable τ with a delta function concentrated on the particle’s trajectory. The resulting expressions then involve the current density j(y) defined in equation (2.5) as well as the energy-momentum tensor (density) of the point particle which is given by (see Appendix A)

Tµν(y) = Z

dτ mx˙µνδ4(y−x(τ)). (2.14)

More explicitly, by using

˙

xννyδ4(y−x(τ)) =−x˙ννxδ4(y−x(τ)) =− d

dτδ4(y−x(τ)), (2.15)

we have

νyTνµ(y) = Z

dτ mx˙µννyδ4(y−x(τ)) =− Z

dτ mx˙µ d

dτδ4(y−x(τ))

= Z

dτ m¨xµδ4(y−x(τ)),

and substitution of the particle’s equation of motion m¨xµ= Tr(qFµν) ˙xν then yields

νyTνµ(y) = Z

dτx˙νqaFaµν(x(τ))δ4(y−x(τ)) =Faµν(y) Z

dτx˙νqaδ4(y−x(τ))

=Faµν(y)jνa(y).

The continuum version of Wong’s equations (2.9) thus reads

νTνµ= Tr(Fµνjν), where Dµjµ= 0. (2.16)

These equations describe the exchange of energy and momentum between the field Fµν and the current jµ (i.e. the matter). They admit an obvious generalization to Moyal space, see equation (3.8) below. In Appendix A, we show that they also admit a natural extension to curved space (endowed with a metric tensor (gµν(x))). Moreover, we will prove there that they have to hold for arbitrary dynamical matter fields φ whose dynamics is described by a generic actionS[φ;gµν, Aaµ] which is invariant under both gauge transformations and general coordinate transformations (gµν and Aaµ representing fixed external fields). We note that the expectation values of equations (2.16) viewed as operatorial relations in quantum field theory imply Wong’s classical equations of motion for sufficiently localized, quantum “wave-packet” states [9].

Curved space. Finally, we also point out that equations which are somewhat similar to Wong’s equations appear in general relativity for a spinning particle in curved space, for which case the contraction of the Riemann tensor Rαβµν with the spin tensorSµν plays a role which is similar to the field strengthFµν in Yang–Mills theories. The explicit form of these equations of motion, which are known as theMathisson–Papapetrou–Dixon equations [13,30,33], is given by

dτ(muα) +1

2SµνuσRασµν = 0, ∇Sαβ dτ = 0,

where denotes the covariant derivative along the trajectory and (uµ) is the particle’s four- velocity.

(7)

3 Lagrangian approach to particles in NC space

Moyal space and distributions. We consider four dimensional Moyal space, i.e. we assume that the space-time coordinates fulfill a Heisenberg-type algebra (for a review see [8,35,39] and references therein). Thus, the star product of functions is defined by

(f ? g)(x)≡ e2iθµνµxνyf(x)g(y) x=y,

where the parameters θµν = −θνµ are constant, and their star commutator is defined by [f ?, g]≡f ? g−g ? f, which implies that [xµ ?, xν] = iθµν. In the sequel we will repeatedly use the following fundamental properties of the star product:

Z

d4xf ? g= Z

d4xf ·g, Z

d4xf ? g ? h= Z

d4xh ? f ? g. (3.1)

An important point to note is that in Moyal space, the integral R

d4xplays the role of a trace4, and hence equations of motion must always be derived from the action rather than the Lag- rangian. For a detailed discussion of the algebras of functions and of distributions on Moyal space in the context of non-commutative spaces and of quantum mechanics in phase space, we refer to [43] and [21, 42], respectively. Here, we only note that the star product of the delta distribution δy (with support in y) with a function ψ may be defined by application to a test functionϕ:

y? ψ, ϕi ≡ Z

d4x(δy? ψ)(x)ϕ(x) = Z

d4x(δy? ψ ? ϕ)(x) = Z

d4y(x)(ψ ? ϕ)(x)

= (ψ ? ϕ)(y).

Hence, the action of the distribution δy? ψ on the test function ϕ is equal to the action of the distributionδy on the test function ψ ? ϕ. Similarly, we find

hψ ? δy, ϕi ≡ Z

d4x(ψ ? δy)(x)ϕ(x) = Z

d4x(ψ ? δy ? ϕ)(x) = Z

d4x(δy? ϕ ? ψ)(x)

= (ϕ ? ψ)(y).

The following considerations hold for an arbitrary antisymmetric matrix (θµν), but for the physical applications it is preferable to assume thatθµ0= 0, i.e. assume the time to be commut- ing with the spatial coordinates. This choice is motivated by the fact that the parameters θij have close analogies with a constant magnetic field both from the algebraic and dynamical points of view [11], and by the fact that a non-commuting time leads to problems with time-ordering in quantum field theory [5].

Charged particle in Moyal space. Since aU?(1) gauge field (Aµ) on Moyal space entails a non-Abelian structure for the field strength tensor (Fµν) due to the star product [8,35,39],

Fµν ≡∂µAν −∂νAµ−ig[Aµ?, Aν],

one expects that the treatment of a source for this gauge field as given by a “point” particle should allow for a description of such particles on non-commutative space which is quite similar to the one found by Wong in the case of Yang–Mills theory. If the matter content in field theory is given by a spinor fieldψ, then the interaction term with the gauge field reads

Z

d4yJµ? Aµ, with Jµ≡gψγ¯ µ? ψ,

4This is best seen by employing the Weyl map from operators to functions, and is also the reason for the cyclicity property (3.1). Furthermore, when considering gauge fields only the action is gauge invariant, not the Lagrangian.

(8)

i.e. it involves two star products. By virtue of the properties (3.1) one of these star products can be dropped under the integral, but not both of them. If we consider the particle limit (i.e. Jµ representing the current density of the particle), it is judicious to maintain the star product betweenJµandAµso as not to hide the non-commutative nature of the underlying space (over which we integrate) and to allow for the use of the cyclic invariance property (3.1) later on in our derivation. In fact, the pairinghJ, Ai ≡R

d4yJµ? Aµ represents the analogue of the pairing hj, Ai ≡ R

d4yTr(jµAµ) in Yang–Mills theory on commutative space. Accordingly, we will require the action for the particle to be invariant under non-commutative gauge transformations defined at the infinitesimal level by δλAµ =Dµλ≡∂µλ−ig[Aµ ?, λ] where the parameter λ is an arbitrary function. As was done for Wong’s equations, we assume that the charge q of the relativistic particle in non-commutative space depends on the parameter τ parametrizing the particle’s world line and that it represents a given non-dynamical variable. The coupling of this particle to an externalU?(1) gauge field (Aµ) can be described by the action5

S[x]≡Sfree[x]−Sint[x]≡ −m Z

˙ x2

Z

d4yJµAµ, (3.2)

where

Jµ(y)≡ Z

dτ q(τ) ˙xµ(τ)δ4(y−x(τ)). (3.3)

Keeping in mind the formulation of field theory on non-commutative space [8, 35, 39], we will argue directly with the action rather than the Lagrangian function and require this ac- tion to be invariant under non-commutative gauge transformations. For such an infinitesimal transformation we have

δλ Z

d4yJµAµ= Z

d4yJµDµλ=− Z

d4y(DµJµ)? λ.

Hence, invariance of the action (3.2) under non-commutative gauge transformations requires the current to be covariantly conserved, i.e. DµJµ = 0. By virtue of equation (3.3) we now infer that

0= (D! µJµ)(y) = Z

dτ qx˙µDyµδ4(y−x(τ))

= Z

dτ qx˙µ

µyδ4(y−x(τ))−ig[Aµ(y)?, δ4(y−x(τ))] . (3.4) Equation (2.15) entails that the first term in the last line can be rewritten as

Z

dτ qx˙µµyδ4(y−x(τ)) =− Z

dτ q d

dτδ4(y−x(τ)) = Z

dτdq

dτδ4(y−x(τ)).

From condition (3.4) it thus follows that the chargeq has to becovariantly conserved along the world line in the sense that

0 = Z

dτDq

dτδ4(y−x(τ))≡ Z

dτ dq

dτδ4(y−x(τ))−igqx˙µ[Aµ(y)?, δ4(y−x(τ))]

. (3.5) For later reference, we note that this relation yields the following equality after star multiplication with Aν(y)δyν and integration overy:

Z d4y

Z

dτ δxν(τ)dq

dτδ4(y−x(τ))? Aν(y)

5We note that the coupling of U?(1) gauge fields to external currents has also been addressed in the recent work [4].

(9)

= ig Z

d4y Z

dτ δxν(τ)qx˙µ[Aµ(y)?, δ4(y−x(τ))]? Aν(y)

=−ig Z

d4y Z

dτ δxν(τ)qx˙µδ4(y−x(τ))?[Aµ(y)?, Aν(y)]. (3.6) In order to derive the equation of motion for the particle determined by the action (3.2), we vary the latter with respect to xµ. The variation of Sfree being the same as in commutative space, we only work out the variation of the interaction partSint, all star products being viewed as functions of the variable y:

δSint =δ Z

d4y Z

˙

xµ4(y−x(τ))? Aµ(y)

= Z

d4y Z

δx˙µ4(y−x(τ))? Aµ(y) + ˙xµ

δ4(y−x(τ))

? Aµ(y)

= Z

d4y Z

d(δxµ)

dτ qδ4(y−x(τ))? Aµ(y) + ˙xµq(δxν)∂νxδ4(y−x(τ))? Aµ(y)

= Z

d4y Z

dτ(δxν)

− d dτ

4(y−x(τ))

? Aν(y) + ˙xµ4(y−x)? ∂νyAµ(y)

. By virtue of the product rule, the first term in the last line yields two terms, one involving

dq

which can be rewritten using relation (3.6), and one involving dδ4(y−x(τ)) which can be rewritten using (2.15):

− Z

d4y Z

dτ(δxν)q d

dτδ4(y−x(τ))? Aν(y)

= Z

d4y Z

dτ(δxν)qx˙µ

µyδ4(y−x(τ))

? Aν(y)

=− Z

d4y Z

dτ(δxν)qx˙µδ4(y−x(τ))? ∂µyAν(y).

Hence, we arrive at δSint =

Z d4y

Z

dτ(δxν) ˙xµ4(y−x(τ))?

νyAµ(y)−∂µyAν(y) + ig[Aµ(y)?, Aν(y)]

= Z

d4y Z

dτ(δxν) ˙xµ4(y−x(τ))? Fνµ(y) = Z

dτ(δxν) ˙xµqFνµ(x(τ)).

Note that it is the constraint equation (3.6) following from (3.5) that yields the terms which are quadratic in the gauge field. This is very much the same mechanism as in the commutative space calculation which leads to Wong’s equations, cf. (2.4)–(2.9).

In conclusion, we obtain the followingequation of motion for the charged relativistic particle in non-commutative space:

mx¨µ=qFµνν, with Fµν ≡∂µAν−∂νAµ−ig[Aµ?, Aν]. (3.7) We note that the antisymmetry ofFµν with respect to its indices implies

0 = ¨xµµ= 1 2

dx˙2 dτ ,

which is consistent with ˙x2 = 1. Consistency of the equation of motion (3.7) requires its gauge invariance, i.e. the gauge invariance of its right-hand-side. Since

δλ(qFµνν) =q(δλFµν) ˙xν =−igq[Fµν ?, λ] ˙xν =gqθρσ(∂ρFµν)(∂σλ) ˙xν +O θ2 ,

(10)

the gauge invariance only holds for constant field strengths by contrast to the case of Wong’s equation for Yang–Mills fields in commutative space. This difference can be traced back to the fact that the analogue of the trace in Yang–Mills theory is given in Moyal space by the integralR

d4x: since such an integral does not occur in the differential equation (3.7), the gauge invariance is only realized for constant Fµν. We will further discuss the latter fields and the comparison with Yang–Mills theory in Section 4 where we will also consider the subsidiary condition for the chargeq. Here we only note the following points in this respect. The equation of motion (3.7) taken for itself is consistent for any constant values of q and Fµν. Furthermore in non-commutative space the auxiliary variable q cannot be rendered dynamical with a non- trivial gauge transformation law (2.10) for q(τ) since the star commutator ofq(τ) with a gauge transformation parameter λ(x(τ)) vanishes.

In summary, the coupling of a relativistic particle to a gauge field (Aµ) is described in general by the Lagrangian

L(x,x) =˙ −m

˙

x2−qAµµ, or L(x,x) =˙ m

2x˙2+qAµµ.

The non-commutativity of space-time can be implemented in the Lagrangian framework by rewriting the interaction term of the action as an integralR

d4y(Jµ?Aµ)(y) where the currentJµ is defined by (3.3) and by requiring this action to be invariant under non-commutative gauge transformations. The resulting equation of motion (3.7) is only gauge invariant for constant field strengths.

Continuum formulation of the dynamics. By following the same lines of arguments as for the Lorentz–Yang–Mills force equation (see equations (2.14)–(2.16)), we can obtain a continuum version of the equation of motion (3.7) by multiplying this equation withδ4(y−x(τ))ϕ(y), where ϕ(y) is a suitable test function, and integrating over τ and over y. More explicitly, by starting from the energy-momentum tensor (2.14) for the point particle and using relation (2.15), we get

Z

d4y(∂νTνµ)(y)ϕ(y) = Z

d4y Z

dτ mx˙µννyδ4(y−x(τ))ϕ(y)

=− Z

d4y Z

dτ mx˙µ d

dτδ4(y−x(τ))ϕ(y) = Z

d4y Z

dτ m¨xµδ4(y−x(τ))ϕ(y).

Substitution of equation (3.7) then yields the expression Z

d4y Z

dτ qFµν(x(τ)) ˙xνδ4(y−x(τ))ϕ(y) = Z

d4y Z

dτ qFµν(y) ˙xνδ4(y−x(τ))ϕ(y)

= Z

d4y(FµνJν)(y)ϕ(y),

where we considered the current density (3.3) in the last line. Thus, we have the result

νTνµ=FµνJν, where DµJµ= 0, (3.8)

and these relations are completely analogous to the continuum equations (2.16) which correspond to Wong’s equations (apart from the fact that the invariance of (3.8) under non-commutative gauge transformations requires the field strength Fµν to be constant).

4 Case of a constant f ield strength

We will now discuss the dynamics of charged particles coupled to a constant field strength on the basis of the results obtained in Sections 2 and 3. Indeed this case represents a mathematically tractable and physically interesting application of the general formalism.

We successively discuss the case of non-Abelian Yang–Mills theory on Minkowski space and the case of a U?(1) gauge field on Moyal space while emphasizing the differences that exist for constant field strengths.

(11)

4.1 Wong’s equations in commutative space

The case of a “uniform field strength” in Yang–Mills theory has been addressed some time ago by the authors of reference [9], see also [19] for related points. Since the Yang–Mills field strength Fµν(x) ≡ Fµνa (x)Ta is not gauge invariant, but transforms under finite gauge transformations as Fµν0 =U−1FµνU (with U(x) ∈G= structure group), one has to specify first what is meant by a constant field. The field Fµν is said to be uniform if the gauge field Aµ(x) ≡ Aaµ(x)Ta at a point x can be related by a gauge transformation to the gauge field Aµ(y) at any other pointy. More precisely, for a space-time translation parametrized bya∈R4, there exists a gauge transformation x7→U(x;a)∈Gsuch that

Aµ(x+a) =U−1(x;a)Aµ(x)U(x;a) + iU−1(x;a)∂µU(x;a), and thereby

Fµν(x+a) =U−1(x;a)Fµν(x)U(x;a).

In this case a gauge may be chosen in which all components of Fµν are constant becauseFµν at the point x can be made equal to its value at some arbitrary point y by transforming it by an appropriate gauge group element.

Since the field strengthFµν ≡∂µAν −∂νAµ+ i[Aµ, Aν] contains two terms, namely ∂µAν

νAµ which has the Abelian form, and [Aµ, Aν] which does not involve derivatives, a constant non-zero field strengthFµν can be obtained either from a linear gauge potential (i.e. an Abelian- like gauge field),

Aµ=−1

2Fµνxν, ∂µAν−∂νAµ=Fµν = const, [Aµ, Aν] = 0, or from a constant non-Abelian-like gauge potential,

Aµ= const, ∂µAν = 0, i[Aµ, Aν] =Fµν = const.

In fact [9], these two types of potentials exhaust all possibilities for a constant field strength.

It has been shown for the structure group SU(2) that the two types of gauge potentials leading to a same constant field strength are gauge inequivalent and result in physically different be- havior when matter interacts with them, e.g. the solutions of Wong’s equations have completely different properties in both cases. An explicit example for G = SU(2) is given by a constant magnetic field inz-direction [19]: letσk(withk= 1,2,3) denote the Pauli matrices and suppose

F0i = 0, FijijkBk, with (Bk)k=1,2,3= (0,0,2σ3).

This constant field strength derives from the linear Abelian-like potential A~ ≡(Ak)k=1,2,3=−1

2~x∧B~ = (−y, x,0)σ3,

or from a constant non-Abelian-like potentialA~= (−σ2, σ1,0).

4.2 Wong’s equations in non-commutative space

The field strength associated to aU?(1) gauge field (Aµ) on Moyal space reads Fµν ≡∂µAν −∂νAµ−ig[Aµ?, Aν] =∂µAν−∂νAµ+gθρσ(∂ρAµ)(∂σAν) +O θ2

.

Due to the derivatives appearing in the star commutator, the Abelian-like term ∂µAν −∂νAµ

and the non-Abelian-like term −ig[Aµ ?, Aν] cannot vanish independently of each other: the

(12)

non-commutative field strength Fµν can only be constant for a linear Abelian-like potential.

More precisely, for Aµ=−1

2B¯µνxν, (4.1)

where the coefficients ¯Bµν ≡ −B¯νµ are constant, we obtain Fµν = ¯Bµν− g

4B¯µρθρσσν. (4.2)

This field strength is constant, but dependent on the non-commutativity parameters θµν. If we interpret ¯Bµν = ∂µAν −∂νAµ as the physical field strength, then (4.2) means that the non- commutativity parametersθµν modify in general the trajectories of the particle as compared to its motion in commutative space6.

Since the field strength transforms under a finite gauge transformation U(x) ∈ U?(1) as Fµν =U−1?Fµν?U, a constant fieldFµνis gauge invariant. Thus, for constant field strengths the situation is quite different forU?(1) gauge fields and for non-Abelian gauge fields on Minkowski space despite the fact that we encounter the same structure for the gauge transformations, the field strength and the action functional in both cases.

Let us again come back to the expression (4.1) for the gauge field. Substitution of this expression into the subsidiary condition (3.5) yields

0 = Z

˙

4(y−x(τ))−igqx˙µ[Aµ(y)?, δ4(y−x(τ))]

= Z

˙

4(y−x(τ))−g

2qx˙µµρθρσσyδ4(y−x(τ)) . (4.3) If the matrix ( ¯Bµν) is the inverse of the matrix (θµν), i.e. ¯Bµρθρσµσ, thenFµν = (1−g4) ¯Bµν and, by virtue of (2.15) and an integration by parts, condition (4.3) takes the form

0 = Z

dτqδ˙ 4(y−x(τ))n 1−g

2 o

.

The latter relation is obviously satisfied for a constant q. In this case, the equation of mo- tion (3.7) for the particle in non-commutative space, i.e. mx¨µ=qFµνν, has the same form as the one of an electrically charged particle in ordinary space. This result is analogous to the one obtained for a constant magnetic field in x3-direction within the Hamiltonian approaches, see equations (5.10) and (5.13) below.

5 Hamiltonian approaches to particles in NC space

To start with, we briefly review the Hamiltonian approaches in commutative space before con- sidering the generalization to the non-commutative setting. In the latter setting, we will notice that various approaches yield different results since several expressions which coincide in com- mutative space no longer agree.

6We note that the latter dependence on the non-commutativity parameters can be eliminated mathematically if one assumes that ¯Bµν depends in a specific way on the parametersθµν and someθ-independent constantsBµν. To illustrate this point [11], we assume that the only non-vanishing components of ¯Bµν and θµν are as follows:

B¯12=B¯21 B, θ¯ 12 =−θ21 θ, i.e.F12= ¯B(1 +g4θB) =¯ −F21. If ¯B depends onθ and on aθ-independent constantB according to

B¯= ¯B(B;θ) 2

p1 +gθB1

=B 1g

4θB

+O θ2 ,

then relation (4.2) implies that the non-commutative field strengthF12 is aθ-independent constant: F12=B.

(13)

5.1 Reminder on the Poisson bracket approach

The Hamiltonian formulation of relativistic (as well as non-relativistic) mechanics is based on two inputs (e.g. see reference [29]): a Hamiltonian function and a Poisson structure (or equivalently a symplectic structure). If one starts from the Lagrangian formulation, the Hamiltonian function is obtained from the Lagrange function by a Legendre transformation. E.g. the Lagrangian L(x,x) =˙ m22+eAµµ (involving the constant chargee) yields theHamiltonian

H(x, p) = 1

2m(p−eA)2 = 1

2m(pµ−eAµ)(pµ−eAµ). (5.1)

The trajectories in phase space are parametrized by τ 7→ (x(τ), p(τ)) where τ denotes a real variable to be identified with proper time after the equations of motion have been derived. The Poisson brackets {·,·} of the phase space variables xµ, pµ are chosen in such a way that the evolution equation ˙F ={F, H} (where ˙F ≡dF/dτ) yields the Lagrangian equation of motion for xµ, though written as a system of first order differential equations. For instance, if we consider the usual form of the Poisson brackets, i.e. the canonical Poisson brackets

{xµ, xν}= 0, {pµ, pν}= 0, {xµ, pν}=ηµν, (5.2) then substitution of F =xµ and F =pµ into ˙F = {F, H} (with H given by (5.1)) yields the system of equations

mx˙µ=pµ−eAµ, mp˙µ=e(pν−eAν)∂µAν, from which we conclude that

mx¨µ= ˙pµ−eA˙µ= e

m(pν −eAν)

| {z }

=ex˙ν

µAν −ex˙ννAµ=ex˙νµAν −∂νAµ

≡efµνν.

This equation coincides with the Euler–Lagrange equation forxµfollowing from the Lagrangian L(x,x) =˙ m22+eAµµ.

Concerning the gauge invariance, we emphasize a result [40] which does not seem to be very well known. The Hamiltonian (5.1) is not invariant under a gauge transformation Aµ → Aµ+∂µλwhich is quite intriguing. However, it is invariant if this transformation is combined with the phase space transformation (xµ, pµ) → (xµ, pµ+e∂µλ), the latter being a canonical transformation since it preserves the fundamental Poisson brackets (5.2). Indeed, under this combined transformation the kinematical momentum Πµ≡pµ−eAµ(which coincides withmx˙µ) is invariant.

We note that the Hamiltonian (5.1) can be rewritten in terms of the variable Πµ≡pµ−eAµ asH≡ 2m1 Π2. TherebyH has the form of a free particle Hamiltonian, but the Poisson brackets are now modified: from Πµ=pµ−eAµ and (5.2) it follows that

{xµ, xν}= 0, {Πµν}=efµν(x), {xµν}=ηµν, (5.3) with fµν ≡∂µAν −∂νAµ. Since the electromagnetic field strength fµν is gauge invariant, the latter invariance is manifestly realized in this formulation.

In summary, the coupling of a charged particle to an electromagnetic field can either be described by the canonical Poisson brackets (5.2) and the minimally coupled Hamiltonian (5.1) or by introducing the field strength into the Poisson brackets (as a non-commutativity of the momenta) and considering a Hamiltonian which has the form of a free particle Hamiltonian.

(14)

5.2 Reminder on the symplectic form approach

If we gather all phase space variables into a vector ξ~ ≡ (ξI) ≡ (x0, . . . , x3, p0, . . . , p3), the fundamental Poisson brackets (5.2) read

ξI, ξJ = ΩIJ, with ΩIJ

0 ηµν

−ηµν 0

.

The inverse of the Poisson matrix Ω ≡(ΩIJ) is the matrix with entries ωIJ ≡(Ω−1)IJ which defines the symplectic 2-form

ω≡ 1 2

X

I,J

ωIJI∧dξJ =dpµ∧dxµ, (5.4)

e.g. see reference [29] for mathematical details. The Hamiltonian equations of motion can be written as

ξ˙I =

ξI, H = ΩKJKξIJH, i.e. ξ˙I= ΩIJJH, or equivalently asωIJξ˙J =∂IH.

In terms of the phase space variables (xµµ) appearing in the non-canonical Poisson brac- kets (5.3), the symplectic 2-form (5.4) reads

ω=dΠµ∧dxµ+1

2efµνdxµ∧dxν.

This formulation of the electromagnetic interaction based on the symplectic 2-form and the evolution equation ωIJξ˙J =∂IH goes back to the seminal work of Souriau [37].

5.3 Standard (Poisson bracket) approach to NC space-time

In order to introduce a non-commutativity for the configuration space, one generally starts from a function H on phase space to which one refers as the Hamiltonian without any ref- erence to a Lagrangian, e.g. we can consider the function H given in equation (5.1). The non-commutativity of space-time is then implemented by virtue of the Poisson brackets

{xµ, xν}=θµν, {pµ, pν}= 0, {xµ, pν}=ηµν, (5.5) whereθµν =−θνµ is again assumed to be constant. (For an overview of the description of non- relativistic charged particles in non-commutative space we refer to [3, 11, 24], the pioneering work being [17,18], see also [1,2,31,32] for some subsequent early work. We also mention that dynamical systems in non-commutative space can be constructed by applying Dirac’s treatment of constrained Hamiltonian systems to an appropriate action functional, see [12] and references therein.)

As in the commutative setting, we gather all phase space variables into a vectorξ~≡(ξI) ≡ (x0, . . . , x3, p0, . . . , p3), the fundamental Poisson brackets (5.5) being now given by

I, ξJ}= ΩIJ, with (ΩIJ)≡

θµν ηµν

−ηµν 0

.

Quite generally, the Poisson bracket of two arbitrary functionsF,Gon phase space reads {F, G} ≡X

I,J

IJIF ∂JG=θµν ∂F

∂xµ

∂G

∂xν + ∂F

∂xµ

∂G

∂pµ − ∂F

∂pµ

∂G

∂xµ.

(15)

Substitution of F = xµ and F = pµ into ˙F = {F, H} (with H given by (5.1)) yields the system of equations

mx˙µ= (pν−eAν)(ηµν−eθµρρAν),

mp˙µ=e(pν −eAν)∂µAν. (5.6)

In the present case, the phase space transformation (xµ, pµ)→(xµ, pµ+e∂µλ) does not represent a canonical transformation since it does not preserve the Poisson brackets (5.5) ifθµν 6= 0. Hence, the resulting Hamiltonian equations of motion (5.6) are not gauge invariant, as has already been pointed out in reference [11] by considering different gauges.

In the next two subsections, we recall how this problem can be overcome for the particular case of a constant field strength as well as more generally, and we compare with the results obtained in Section 3 from the action involving star products. Here, we only note that a non- Abelian structure of the field strength is hidden in equation (5.6). To illustrate this point, we consider the particular case where the only non-vanishing components of θµν are θij = εijθ (with i, j ∈ {1,2} and ε12 ≡ −ε21 ≡ 1) and where the only non-vanishing components of Aµ areAi(x1, x2) (withi∈ {1,2}). For this situation which describes a time-independent magnetic field perpendicular to thex1x2-plane, the first of equations (5.6) yields

mx˙i = (pk−eAk)(δik−eθεijjAk), and implies

md

dt(xi+eθεijAj) = (1 +eF12)(pi−eAi) (5.7) with

F12≡∂1A2−∂2A1+e{A1, A2}=∂1A2−∂2A1+eθρσ(∂ρA1)(∂σA2). (5.8) Thus, we find a non-Abelian structure for the generalized field strength, but in the present approach the field Fµν is only linear in the non-commutativity parameters in contrast to the fieldFµν in (3.7) which involves the star commutator

−i[Aµ?, Aν] =θρσ(∂ρAµ)(∂σAν) +O θ2 .

If the gauge potential is linear inx, the field strengthsFµνandFµνas defined by equations (5.8) and (3.7), respectively, coincide with each other (if one identifies the coupling constantgwithe).

To conclude, we note that (5.7) can be solved for pi−eAi in terms of mx˙i: the system of first order differential equations (5.6) can then be written as a second order equation for xµ, but the resulting equations of motion are not gauge invariant and they cannot be derived from a Lagrangian [3].

5.4 Standard approach to NC space-time continued

The reasoning presented concerning the brackets (5.3) suggests to consider a Hamiltonian which has a free form and to introduce a field strengthBµν(x) as a non-commutativity of the momenta, i.e. consider phase space variables (xµ, pµ) satisfying the non-canonical Poisson algebra

{xµ, xν}=θµν, {pµ, pν}=eBµν, {xµ, pν}=ηµν, (5.9) with θµν constant. As pointed out in reference [22], the Jacobi identities for the algebra (5.9) are only satisfied if the field strength is constant:

{xµ,{pν, pλ}}+{cyclic permutations of µ,ν,λ}=eθµρρBνλ.

参照

関連したドキュメント

Finite difference operator on words Non commutative Gandhi polynomials The Dumont-Foata polynomials. Commutative version Non commutative version A combinatorial

Merle; Global wellposedness, scattering and blow up for the energy critical, focusing, nonlinear Schr¨ odinger equation in the radial case, Invent.. Strauss; Time decay for

Later, in [1], the research proceeded with the asymptotic behavior of solutions of the incompressible 2D Euler equations on a bounded domain with a finite num- ber of holes,

“Breuil-M´ezard conjecture and modularity lifting for potentially semistable deformations after

Then it follows immediately from a suitable version of “Hensel’s Lemma” [cf., e.g., the argument of [4], Lemma 2.1] that S may be obtained, as the notation suggests, as the m A

[9, 28, 38] established a Hodge- type decomposition of variable exponent Lebesgue spaces of Clifford-valued func- tions with applications to the Stokes equations, the

Any nonstandard area-minimizing double bubble in H n in which at least one of the enclosed regions is connected consists of a topological sphere intersecting the axis of symmetry

Any nonstandard area-minimizing double bubble in H n in which at least one of the enclosed regions is connected consists of a topological sphere intersecting the axis of symmetry